Highly Enantioselective Chlorination of β-Keto Esters and Subsequent

May 31, 2012 - Construction of Chiral α-Amino Quaternary Stereogenic Centers via Phase-Transfer Catalyzed Enantioselective α-Alkylation of α-Amidom...
0 downloads 11 Views 829KB Size
Communication pubs.acs.org/JACS

Highly Enantioselective Chlorination of β-Keto Esters and Subsequent SN2 Displacement of Tertiary Chlorides: A Flexible Method for the Construction of Quaternary Stereogenic Centers Kazutaka Shibatomi,* Yoshinori Soga, Akira Narayama, Ikuhide Fujisawa, and Seiji Iwasa Department of Environmental and Life Sciences, Toyohashi University of Technology, 1-1 Hibarigaoka, Tempaku-cho, Toyohashi 441-8580, Japan S Supporting Information *

Scheme 1

ABSTRACT: Highly enantioselective chlorination of βoxo esters and subsequent stereospecific substitution of tertiary chlorides are described. Enantioselective chlorination of β-keto esters and malonates was performed using a chiral Lewis acid catalyst prepared from Cu(OTf)2 and the newly developed spirooxazoline ligand 2 to yield the desired α-chlorinated products with high enantioselectivity (up to 98% ee). Nucleophilic substitution of the resulting chlorides proceeded smoothly to afford a variety of chiral molecules such as α-amino, α-alkylthio, and α-fluoro esters, without loss of enantiopurity. The results of X-ray crystallographic analysis proved that Walden inversion occurs at the chlorinated tertiary carbon center. These results supported the fact that the substitution proceeds via an SN2 mechanism.

We first applied SPYMOX to the asymmetric chlorination of βketo esters using N-chlorosuccinimide (NCS) as the chlorination reagent. Chlorination of 3a in the presence of the SPYMOX/Cu(OTf)2 complex afforded the desired product 4a in high yield but with moderate enantioselectivity (78% ee; Table 1, entry 1). To enhance the enantioselectivity in this Table 1. Optimization of the Reaction Conditionsa

T

he SN2 reaction, one of the most fundamental reactions in organic transformation, can be used for the construction of enantioenriched chiral carbon centers because the substitution proceeds with complete inversion of stereochemistry. Optically active tertiary chlorides could be useful intermediates for the preparation of a variety of chiral molecules having a quaternary stereogenic center if SN2 substitution were to occur at the chlorinated tertiary carbon. However, this process has not been successful to date except in a rare instance1 because of two major underlying problems: First, as described in organic chemistry textbooks, tertiary halides (particularly tertiary chlorides) rarely undergo SN2 substitution.2 Second, there are relatively few catalytic methods that afford highly enantioenriched tertiary chlorides, and the development of such a reaction with broad substrate scope remains a challenge.3−5 We describe herein the highly enantioselective chlorination of a wide range of β-keto esters in the presence of a newly synthesized chiral spirooxazoline ligand and subsequent unimpeded SN2 substitution at the chlorinated tertiary carbon. The present method yields a variety of chiral molecules having a quaternary stereogenic center (Scheme 1). We began by focusing on the development of an efficient catalyst for the enantioselective chlorination. Our recent research revealed that the chiral pyridyl spirooxazoline ligand 1, which we call SPYMOX, shows high asymmetric induction ability in the palladium-catalyzed allylic alkylation6a and copper(II)-catalyzed asymmetric fluorination of β-keto esters.6b © XXXX American Chemical Society

entry

ligand

solvent

time (h)

% yieldb

% eec

1 2 3 4 5 6

1 2a 2b 2a 2a 2a

benzene benzene benzene toluene CH2Cl2 THF

1 1 1 1 1 2

93 97 96 96 93 91

78 95 95 94 93 71

a

All of the reactions were carried out using 1.5 equiv of NCS in the presence of a Lewis acid catalyst prepared from 12 mol % chiral ligand and 10 mol % Cu(OTf)2. bIsolated yields. cDetermined by chiral GC analysis.

chlorination, we designed new chiral spirooxazoline ligands 2 with a quinoline backbone. Specifically, we expected the steric repulsion between the substrate and the additional benzene Received: January 26, 2012

A

dx.doi.org/10.1021/ja304806j | J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Journal of the American Chemical Society

Communication

ring in 2 to push the reaction center closer to the chiral binaphthyl backbone. As expected, the enantioselectivity of the reaction dramatically increased to 95% ee when 2 was used under the same reaction conditions (entries 2 and 3). Screening of various solvents showed that benzene afforded the best enantioselectivity and reactivity. With the optimized reaction conditions, we attempted to expand the substrate scope of the asymmetric chlorination. As summarized in Table 2, various cyclic and acyclic β-keto esters, including both aliphatic and aromatic keto esters, were successfully chlorinated with high enantiopurity. It was found that a bulky ester substituent is essential for obtaining high selectivity (entry 6 vs entry 7). Chlorination of αalkylmalonates also proceeded with good-to-high enantioselectivity (entries 19−22).7 On the other hand, chlorination of αcyanoacetate and β-ketophosphonates gave poor enantioselectivity (entries 23−25). In the chlorination process, certain functional groups such as carbamate, alkene, nitrile, and an additional ketone were well-tolerated (entries 4, 5, 13−15, 21, and 22). The sense of stereoselection in the chlorination was controlled by the chiral ligand even when α-chloro-β-keto esters 4i and 4j, which have a menthyl ester as a chiral auxiliary, were used in the reaction (entries 9 and 10). The absolute configuration of the chlorinated chiral carbon in 4j was determined to be S by X-ray structural analysis (see the Supporting Information). Encouraged by the high asymmetric induction ability of the new chiral ligands, we next focused on the SN2 substitution of α-chloro-β-oxo esters.8 First, we attempted to use sodium azide as the nucleophile. The resulting azide could be converted into a primary amine by subsequent palladium-catalyzed hydrogenolysis. For all of the substrates used, the reaction proceeded smoothly to yield the corresponding α-azido esters 5 and αamino esters 6 in high yield (Table 3). Notably, when the palladium loading was increased to 20 mol % in the hydrogenation of 5f and 5h, reduction of the ketone carbonyl proceeded simultaneously with the hydrogenolysis to yield βhydroxy-α-amino esters 7f and 7h in a highly diastereoselective manner (96−99% de; entries 5 and 6).9 Most importantly, the enantiopurity of 5 was exactly the same as that of the starting compound 4 in all cases. Furthermore, single-crystal X-ray crystallographic analysis revealed that the nucleophilic substitution of 4 to give 5 involved a Walden inversion (see the Supporting Information for details). These results strongly suggested that this nucleophilic substitution proceeds via a rigorous SN2 mechanism. The electron-withdrawing nature of the two carbonyl groups next to the chlorinated tertiary carbon probably enables the unusual SN2 substitution of the tertiary chloride.2b Thus, this method will be useful for the preparation of highly substituted α-amino acid derivatives. We next used alkylthiols as nucleophiles. As shown in Table 4, nucleophilic substitution of α-chloro-β-keto esters proceeded smoothly in the presence of triethylamine to yield the corresponding α-alkylthio-β-keto esters 8,10 and the enantiopurity was maintained in all cases (entries 1−7). On the other hand, the reaction of α-chloromalonate 4t with benzylthiol did not afford the desired product at all. Finally, we attempted to carry out a much more challenging reaction, nucleophilic substitution with fluorides for the construction of fluorinated stereogenic centers. After screening several fluorides such as NaF, KF, and tetrabutylammonium fluoride, we found that a combination of CsF and 18-crown-6 efficiently promotes the nucleophilic fluorination of 4f to afford the desired α-fluoro-β-

Table 2. Enantioselective Chlorination of Active Methine Compoundsa

a

All of the reactions were carried out using 1.5 equiv of NCS in the presence of a Lewis acid catalyst prepared from 12 mol % 2a and 10 mol % Cu(OTf)2 , unless otherwise noted. b Isolated yields. c Determined by chiral HPLC or GC analysis. dLigand 2b was used instead of 2a. eDiastereomeric excess. fAbsolute configuration of the chlorinated carbon. gThe reaction was carried out for 48−60 h. h36 mol % 2 and 30 mol % Cu(OTf)2 were used. iThe reaction was carried out under reflux conditions for 24 h with 12 mol % 2b and 10 mol % Cu(OTf)2. jThe reaction was carried out under reflux conditions.

keto ester 9f in high yield with complete Walden inversion (entry 8). The reaction of 4h also gave the desired product 9h without loss of enantiopurity, but the yield was disappointingly low (35%), and tert-butyl 1-hydroxy-2-naphthoate, which was probably generated by the elimination of hydrogen chloride and subsequent aromatization,11 was obtained as a byproduct in B

dx.doi.org/10.1021/ja304806j | J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Journal of the American Chemical Society

Communication

Table 3. SN2 Substitution with Sodium Azide and Subsequent Hydrogenolysis To Give α-Amino Acid Derivativesa

Table 4. SN2 Reaction with Alkylthiols and Cesium Fluoridea

a

All of the reactions were carried out using 3 equiv of NaN3. bIsolated yields of amine 6 or 7 for two steps. Yields in parentheses are the isolated yields of azide 5. cEnantiomeric excess of 5 determined by chiral HPLC or GC analysis. dAzidation was carried out at 25 °C for 6−20 h with 20 mol % Pd catalyst. e96% de. f99% de.

a

Sulfenylation was carried out using 3 equiv of alkylthiol and 6 equiv of Et3N in CH2Cl2. Fluorination was carried out using 3 equiv of CsF and 1 equiv of 18-crown-6 in acetonitrile, unless otherwise noted. b Isolated yields. cDetermined by chiral HPLC analysis. dThe reaction was carried out at 0 °C. eThe reaction was carried out at 80 °C for 24 h. fReaction was carried out at 0 °C for 168 h with 6 equiv of CsF and 1 equiv of 18-crown-6. gThe yield in parentheses was determined by 1 H NMR analysis using 2,2,2-trifluoroethanol as an internal standard. The isolated yield was low because of the volatility of the product.

45% yield (entry 9). Surprisingly, fluorination of 4k (95% ee) yielded the corresponding product 9k with decreased enantiopurity (84% ee; entry 10),12 probably because of the competitive SN1 pathway. Furthermore, the enantiopurity of the starting chloride decreased during the course of the reaction (7% of 4k was recovered, with 84% ee). This partial racemization, which may be due to the nucleophilic substitution (SN2 and/or SN1) of 4k by CsCl generated in situ, was also the reason for the isolation of 9k with lower enantiopurity. Unfortunately, fluorination of other substrates (4a, 4b, 4p, 4r, and 4t) afforded only trace amounts of the desired product under similar reaction conditions, and significant amounts of byproduct were formed. Nevertheless, an important feature of this fluorination method is that a highly enantioenriched fluorinated stereogenic center can be constructed by using the combination of an inexpensive oxidant (NCS) and a nucleophilic fluorination reagent (CsF) instead of a commonly used electrophilic fluorinating reagent13 such as N-fluorobenzenesulfonimide. In conclusion, we have successfully carried out the highly enantioselective α-chlorination of β-keto esters and malonates with a chiral spirooxazoline ligand. Substitution of the resulting chlorides with some nucleophiles successfully afforded the corresponding α-heteroatom-substituted β-oxo esters without loss of enantiopurity. The results of X-ray crystallographic analysis proved that the nucleophilic substitution with sodium azide involved a Walden inversion. These results strongly

suggest that the substitution proceeds via the SN2 mechanism, although the reaction occurred at a tertiary carbon. An advantage of our method is that it allows for the formation of multiple optically active compounds with a quaternary stereogenic center from a single intermediate.



ASSOCIATED CONTENT

S Supporting Information *

Experimental details, characterization data, HPLC and GC enantiomer analyses, and NMR spectra (1H, 13C, and 19F) for new compounds. This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Author

[email protected] Notes

The authors declare no competing financial interest. C

dx.doi.org/10.1021/ja304806j | J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Journal of the American Chemical Society



Communication

(10) For examples of enantioselective electrophilic sulfenylation of βketo esters, see: (a) Ishimaru, T.; Ogawa, S.; Tokunaga, E.; Nakamura, S.; Shibata, N. J. Fluorine Chem. 2009, 130, 1049−1053. (b) Fang, L.; Lin, A.; Hu, H.; Zhu, C. Chem.Eur. J. 2009, 15, 7039−7043. (c) Jereb, M.; Togni, A. Chem.Eur. J. 2007, 13, 9384−9392. (d) Srisailam, S. K.; Togni, A. Tetrahedron: Asymmetry 2006, 17, 2603−2607. (e) Sobhani, S.; Fielenbach, D.; Marigo, M.; Wabnitz, T. C.; Jørgensen, K. A. Chem.Eur. J. 2005, 11, 5689−5694. (f) Jereb, M.; Togni, A. Org. Lett. 2005, 7, 4041−4043. (11) (a) Cui, L.-Q.; Dong, Z.-L.; Liu, K.; Zhang, C. Org. Lett. 2011, 13, 6488−6491. (b) Hart, H. Chem. Rev. 1979, 79, 515−528. (12) The isolated yield of 9k was modest because of its volatility. We confirmed that the enantiopurity of 9k did not change during the solvent evaporation. For the self-disproportionation effect of the enantiomers, see: (a) Ueki, H.; Yasumoto, M.; Soloshonok, V. A. Tetrahedron: Asymmetry 2010, 21, 1396−1400. (b) Soloshonok, V. A. Angew. Chem., Int. Ed. 2006, 45, 766−769. (c) Soloshonok, V. A.; Ueki, H.; Yasumoto, M.; Mekala, S.; Hirschi, J. S.; Singleton, D. A. J. Am. Chem. Soc. 2007, 129, 12112−12113. (13) For recent reviews of asymmetric fluorination, see: (a) Lectard, S.; Hamashima, Y.; Sodeoka, M. Adv. Synth. Catal. 2010, 352, 2708− 2732. (b) Ueda, M.; Kano, T.; Maruoka, K. Org. Biomol. Chem. 2009, 7, 2005−2012. (c) Ma, J.-A.; Cahard, D. Chem. Rev. 2008, 108, PR1− PR43. (d) Furuya, T.; Kuttruff, C. A.; Ritter, T. Curr. Opin. Drug Discovery Dev. 2008, 11, 803−819. (e) Brunet, V. A.; O’Hagan, D. Angew. Chem., Int. Ed. 2008, 47, 1179−1182. (f) Shibata, N.; Ishimaru, T.; Nakamura, S.; Toru, T. J. Fluorine Chem. 2007, 128, 469−483.

ACKNOWLEDGMENTS This study was supported by the Asahi Glass Foundation and a Grant-in-Aid for Young Scientists (B) (23750111) from MEXT. We thank the Instrument Center of the Institute for Molecular Science for permission and advice on the usage of their X-ray diffractometer. We also thank the Nippon Synthetic Chemical Industry Co., Ltd. for supplying ethyl isocyanoacetate used in the synthesis of chiral ligands.



REFERENCES

(1) The following paper reports that the nucleophilic substitution of 2-chloro-2-phenylpropanal by sodium azide proceeds via SN2 displacement: Masaki, Y.; Arasaki, H.; Iwata, M. Chem. Lett. 2003, 32, 4−5. (2) For the SN2 reaction of tertiary halides and related studies, see: (a) Mascal, M.; Hafezi, N.; Toney, M. D. J. Am. Chem. Soc. 2010, 132, 10662−10664. (b) Edwards, O. E.; Grieco, C. Can. J. Chem. 1974, 52, 3561−3562. (c) Grob, C. A.; Seckinger, K.; Tam, S. W.; Traber, R. Tetrahedron Lett. 1973, 14, 3051−3054. (d) Cook, D.; Parker, A. J. J. Chem. Soc. B 1968, 142−148. (e) Miotti, U.; Fava, A. J. Am. Chem. Soc. 1966, 88, 4274−4275. (f) Winstein, S.; Smith, S.; Darwish, D. Tetrahedron Lett. 1959, 1, 24−31. (g) Hughes, E. D.; Ingold, C. K.; Mackie, J. D. H. J. Chem. Soc. 1955, 3173−3177. (3) Although the electrophilic chlorination of β-keto esters is known to be a useful method for the preparation of chiral tertiary chlorides, there is no known catalyst system that gives greater than 90% ee for both acyclic and cyclic substrates. For earlier reports of enantioselective chlorination of β-keto esters, see: (a) Jiang, J. J.; Huang, J.; Wang, D.; Yuan, Z.-L.; Zhao, M.-X.; Wang, F.-J.; Shi, M. Chirality 2011, 23, 272−276. (b) Hintermann, L.; Perseghini, M.; Togni, A. Beilstein J. Org. Chem. 2011, 7, 1421−1435. (c) Cai, Y.; Wang, W.; Shen, K.; Wang, J.; Hu, X.; Lin, L.; Liu, X.; Feng, X. Chem. Commun. 2010, 1250−1252. (d) Etayo, P.; Badorrey, R.; Díaz-de-Villegas, M. D.; Gálvez, J. A. Adv. Synth. Catal. 2010, 352, 3329−3338. (e) Kawatsura, M.; Hayashi, S.; Komatsu, Y.; Hayase, S.; Itoh, T. Chem. Lett. 2010, 39, 466−467. (f) Qi, M.-H.; Wang, F.-J.; Shi, M. Tetrahedron: Asymmetry 2010, 21, 247−253. (g) Frings, M.; Bolm, C. Eur. J. Org. Chem. 2009, 4085−4090. (h) Shibatomi, K.; Tsuzuki, Y.; Iwasa, S. Chem. Lett. 2008, 37, 1098−1099. (i) Bartoli, G.; Bosco, M.; Carlone, A.; Locatelli, M.; Melchiorre, P.; Sambri, L. Angew. Chem., Int. Ed. 2005, 44, 6219−6222. (j) Shibata, N.; Kohno, J.; Takai, K.; Ishimaru, T.; Nakamura, S.; Toru, T.; Kanemasa, S. Angew. Chem., Int. Ed. 2005, 44, 4204−4207. (k) Marigo, M.; Kumaragurubaran, N.; Jørgensen, K. A. Chem.Eur. J. 2004, 10, 2133−2137. (l) Ibrahim, H.; Kleinbeck, F.; Togni, A. Helv. Chim. Acta 2004, 87, 605−610. (m) Hintermann, L.; Togni, A. Helv. Chim. Acta 2000, 83, 2425−2435. (4) For enantioselective chlorination of 3-substituted oxindoles, see: (a) Zheng, W.; Zhang, Z.; Kaplan, M. J.; Antilla, J. C. J. Am. Chem. Soc. 2011, 133, 3339−3341. (b) Zhao, M.-X.; Zhang, Z.-W.; Chen, M.-X.; Tang, W.-H.; Shi, M. Eur. J. Org. Chem. 2011, 3001−3008. (c) Wang, D.; Jiang, J.-J.; Zhang, R.; Shi, M. Tetrahedron: Asymmetry 2011, 22, 1133−1141. (5) For reviews of the enantioselective construction of chlorinated chiral stereogenic centers, see: (a) Shibatomi, K. Synthesis 2010, 2679−2702. (b) France, S.; Weatherwax, A.; Lectka, T. Eur. J. Org. Chem. 2005, 475−479. (6) (a) Shibatomi, K.; Muto, T.; Sumikawa, Y.; Narayama, A.; Iwasa, S. Synlett 2009, 241−244. (b) Shibatomi, K.; Narayama, A.; Soga, Y.; Muto, T.; Iwasa, S. Org. Lett. 2011, 13, 2944−2947. (7) For the Lewis acid-catalyzed enantioselective α-fluorination of malonates, see: Reddy, D. S.; Shibata, N.; Nagai, J.; Nakamura, S.; Toru, T.; Kanemasa, S. Angew. Chem., Int. Ed. 2008, 47, 164−168. (8) We recently disclosed that the nucleophilic substitution of αchloro-α-fluoro-β-keto esters proceeds in a stereospecific manner (see ref 6b). This result encouraged us to try the SN2 displacement of tertiary chlorides. (9) The relative configuration of 7h was determined by X-ray structural analysis. See the Supporting Information for the details. D

dx.doi.org/10.1021/ja304806j | J. Am. Chem. Soc. XXXX, XXX, XXX−XXX